A comprehensive characterization of lattice Boltzmann (LB) schemes to perform warm fluid numerical simulations of particle wakefield acceleration (PWFA) processes is discussed in this paper. The LB schemes we develop hinge on the moment matching procedure, allowing the fluid description of a warm relativistic plasma wake generated by a driver pulse propagating in a neutral plasma. We focus on fluid models equations resulting from two popular closure assumptions of the relativistic kinetic equations, i.e., the local equilibrium and the warm plasma closure assumptions. The developed LB schemes can, thus, be used to disclose insights on the quantitative differences between the two closure approaches in the dynamics of PWFA processes. Comparisons between the proposed schemes and available analytical results are extensively addressed.
I. INTRODUCTION
The process of particle acceleration plays a role of primary importance at the interface between fundamental and applied physics.1 The growing costs and dimensions of conventional large-scale accelerators demand for new and more efficient technologies to push the energies reached by particle beams beyond the state of the art of modern day capabilities. In this context, plasma acceleration is a promising technique that would enable the construction of compact particle accelerators while retaining the same (or superior) energy gains obtained with conventional methods.2–4 A ionized gas is perturbed via the injection of relativistic charged particles (particle wakefield acceleration, PWFA)5 or intense lasers (laser wakefield acceleration, LWFA),6 generally named driver: the interaction between the neutral plasma and the injected driver creates a wave like dynamics of positive and negative charges, and hence, strong accelerating fields (up to 100 GV/m) are developed; the interested reader might look into7,8 to go into detail on the topic. The described process involves a large number of “actors”: the injected driver components, whether particles moving near the speed of light—typically electrons—or laser fields, and both the ions and electrons that make up the plasma. All of them interact with each other via electromagnetic forces, thus making it really difficult to predict and control the final behavior of a plasma acceleration experiment. Theoretical modeling and numerical simulations are, therefore, a powerful tool to help guide the design of new experiments.
On the side of numerical simulations, the most commonly used techniques in the field are represented by particle in cell (PIC) methods9–13 that employ single particle dynamics to describe both the particles in the driver (in this paper, we will focus on PWFA) and the plasma components. These schemes are deeply fine-grained and, while they can capture the most refined phenomena that happen at the microscopic scale, they bring in low-statistics numerical noise.9,14 An alternative numerical modeling can be proposed by recurring to continuum fluid descriptions of the system. Fluid models describe the plasma via macroscopic fields such as particle number density and fluid velocity, and they do so by solving the inviscid relativistic Euler equations that can be systematically derived from kinetic theory of charged gases via a coarse graining of the relativistic Maxwell–Vlasov system15–17 (in their most general warm formulation, these equations are not closed and, hence, require additional constrains—more on this later on). Despite losing the ability to describe some kinetically pertinent features, numerical methods relying on the fluid description are still able to capture non trivial features of the PWFA system and are set up by construction not to show statistical noise. Some examples of fluid solvers used in the realm of PWFA are Architect,18 QFLUID,19 MARPLE,20 FLASH,21 and the code used for hydrodynamic optically field-ionized plasma channels in Ref. 22.
From the theoretical point of view, fluid models have traditionally been developed by neglecting thermal effects, i.e., by neglecting pressure terms in the Euler equations. This choice has been motivated first and foremost by the fact that initial electron thermal energy in the plasma is expected to be of the order of eV,23,24 which is a small value when compared with the electron rest energy MeV (the initial thermal energy normalized to the electron rest energy, , is the control parameter that is usually used to assess the importance of thermal effects), and second by the fact that these cold fluid models are easier to treat theoretically, providing even analytical results in some simplified cases.25–27
Nevertheless, there is a series of important reasons that drive the development of warm fluid models for PWFA. First, the wave-like solutions to cold fluid equations become singular in the presence of wide and highly charged driver pulses (Wave Breaking28,29) The presence of thermal effects may be one of the regularizing mechanisms that mitigate the singularity,30–33 by altering the wakefield properties and allowing electron trapping in the wakefield.34 A significant heating is also expected in the post-wavebreaking dynamic.35 Second, studies targeted at the late stage dynamics of the process36–40 point to the importance of electrons temperature (with particular emphasis on thermally driven ion-acoustic motion39–43) for the restoration of the equilibrium conditions after a driver pulse has passed in the plasma channel. Understanding the restoration conditions is pivotal to enable the possibility of having high repetition rates of driver pulses in order to create sustained accelerating fields. Third, although the aforementioned initial temperatures would not lead to meaningful divergences in behavior from the cold case (at least in the first wave periods34,35), a different situation is expected as many consecutive pulses are injected into the system. The energy deposited by every pulse would be partially transferred to the plasma,36,38 leading to significant increases in temperature [some estimates provide O(1) keV increases in post wave-breaking situations39,40]. Finally, we mention that temperature effects might also be relevant for the study of positron acceleration in quasi-hollow warm plasma channels.41,44–47
If one wants to develop a warm fluid theory, often the only viable option when trying to derive analytical results, an immediate problem has to be tackled: additional fields are now present in the set of equations (namely, the pressure tensor fields), and therefore, a suitable closure has to be carefully selected. The closure problem has been studied in the community, and various models have been proposed48–53 (see Ref. 54 for an outlook).
In this paper, we explore two closures. The first one48 is based on the assumption that the underlying probability distribution solving for the Vlasov equation is a Maxwellian equilibrium, which is described in the relativistic framework by a Maxwell–Jüttner distribution.55 Hereafter, we will refer to it as the local equilibrium closure (LEC). This choice leads to no entropy production and, therefore, grants the adoption of an isentropic equation of state to close for the pressure tensor field that in this framework can be described as a single scalar quantity. The LEC is in principle not well suited for the description of early stage dynamics, as the restoring mechanism that drives the plasma back to an eventual initial equilibrium state (particle collisions) happens on longer time scales than the ones which are typical of the early phases of PWFA. Nevertheless, this closure would still retain a physical relevance when studying late stage dynamics (when particle collisions start to become relevant40). Furthermore, side-by-side comparisons against other closure schemes (or PIC solvers) might, indeed, reveal that the LEC is also helpful for qualitative assessments on early dynamics.
The second closure (hereafter named warm closure—WARMC) is based on the idea proposed in Refs. 49, 50, and 52 and later on reconsidered by Refs. 32, 56, and 57: the centralized moments equations obtained from the coarse-graining of the Vlasov equation are closed by neglecting the third order centralized moment, choice motivated by the assumption of weakly warm systems, and hence small momentum distribution variances. This leads to a closed set of equations that can be solved without having to make hypotheses on the underlying momentum probability distribution, except for the smallness of its variance. This gives also the possibility to evolve independently the various components of the pressure tensor, and hence to evaluate the expected momentum spread anisotropies: in fact, as the dynamics of the system is strongly focused in the acceleration direction, and no collisions are taken into account on these short timescales to regularize the process, momentum spread anisotropies are to be expected.58,59
The main target of this paper is to develop novel numerical schemes for the simulation of warm fluid models in the context of PWFA, for both the LEC and the WARMC. The schemes rely on the lattice Boltzmann (LB) method60,61 to solve the fluid equations. LB is a popular numerical scheme commonly used in computational fluid dynamics as an alternate scheme to direct hydrodynamical solvers. Its formulation is rooted in the kinetic theory of gases, and this provides a strong physical basis to the method. Furthermore, the space locality of the calculations involved in the method makes this solver prone to multi CPUs and/or multi GPUs parallelization.60 In the past years, LB has been generalized to work in many fields other than classical fluid dynamics,61 and it is now widely accepted as a numerical solver for many physics problems that rely on a set of continuum equations. The LB formulation used in this paper is the so-called moment matching LB that solves systems described by advection diffusion equations.60,61 This is the very same formulation used in Ref. 62 to solve for the cold fluid models in PWFA. Here, we extend such formulation to warm fluid models; hence, we are pushing further the usage of the LB method in the context of PWFA. The LB schemes for warm fluids are coupled to a finite difference time domain (FDTD) scheme that solves for the electromagnetic fields:63,64 we refer to Ref. 62 for a more in depth explanation of the coupling. The development of the LB schemes for two different warm fluids closures will enable us to look for thermal effects that depend on the adopted closure scheme and, furthermore, to assess the importance of thermal spreads anisotropies showing side-by-side comparisons between the two closures.
This paper is organized as follows: in Sec. II, we provide a basic introduction to the LB method, with particular emphasis on the moment matching procedure; in Sec. III, basic equations for collisionless relativistic plasmas are reviewed, and the two fluid closures LEC and WARMC are discussed in Secs. IV and V, respectively; numerical results will be presented in Sec. VI, and an outlook and a discussion on future perspectives are given in Sec. VII.
II. LATTICE BOLTZMANN (LB) METHOD
In this section, we introduce the lattice Boltzmann (LB) method, which we use to solve the fluid equations both in the LEC and WARMC models. We first present the basics of the method, that are directly drawn from the kinetic theory of gases and in their original formulation tasked to the reproduction of classical (non relativistic) Navier–Stokes equations. We then illustrate how the whole procedure can be adapted to the simulation of generic advection equations (moment matching LB). We will see in the following sections how the relativistic warm fluid equations, in both the LEC (Sec. IV) and WARMC (Sec. V), can be recast into a set of advection equations and, hence, can be numerically solved via moment matching LB. The interested reader might look into60,61 for more detailed discussions on both LB in its original formulation and its moment matching variant.
A. Basic LB
It can then be verified through the Chapman–Enskog expansion60,61,66 that the moments obtained according to Eq. (5) verify the target continuum equations—again, the Navier–Stokes equations in this original formulation of the LB method—provided that f is sufficiently close to f eq. The most pivotal step in the development of the LB method is the realization that one needs only a truncated version of the distribution functions f and f eq to properly recover the desired moments that solve the field equations.67,68 For this reason, it proves expedient to expand both f and f eq into series of orthogonal Hermite polynomials in the variable and then to truncate this expansion up to the point where one recovers the macroscopic observables of interest.69
The algorithmic steps of the LB scheme are now clearly outlined. A set of Npop versions of Eq. (7), one for every fi, is evolved on a regular spatial grid: the fi are updated at every node with the source term Eq. (8) (Source step) and then stream to neighboring nodes with their corresponding velocity (Streaming step). At every iteration, the hydrodynamic fields are obtained through the discrete summation appearing in Eq. (6).
B. Moment matching procedure for forced advection diffusion equation
III. RELATIVISTIC KINETIC EQUATIONS FOR COLLISIONLESS PLASMA
In this section, we review the basic equations that can be used to build hydrodynamic models of warm plasmas starting from the kinetic theory of gases, all expressed within the framework of special relativity.15–17 In Sec. IV and Sec. V, we will see how the obtained set of equations can be closed via either a local equilibrium assumption or a warm closure and then explain how to recast them into a set of advection equations.
In the following, we will work in a flat space-time with Minkowski metric signature . When expressing formulas in a manifestly covariant form, we will adopt Einstein's summation convention, with Greek indexes running from 0 to and Latin indexes from 1 to . When not explicitly stated, we will use .
IV. LOCAL EQUILIBRIUM CLOSURE (LEC)
A. Local equilibrium closure lattice Boltzmann (LEC-LB)
The two-way coupling of the fluid with the electromagnetic fields needs no particular explanation: the plasma hydrodynamic quantities (particle density n and the transport velocity u) are obtained from the LB evolved quantities (the various components of A) and then fed to the FDTD Maxwell solver together with the driving bunch terms. The evolved electromagnetic fields are then plugged into Eq. (34) as source terms, and the next LB iteration can be performed.
The only detail worth of discussion is the determination of the transport velocity u from the LB-advected quantities A. In fact, due to the appearance of the γ Lorentz factor in Eq. (33), one cannot easily obtain the relationship between u and the second component of the vector A (the one containing momentum , and therefore, there is the need for a specific iterative algorithm to determine it. Initially, set as the zeroth order moment of the distribution function coming out of the LB iteration, divided by n: Then,
- Starting from k = 0, compute as
- Compute as
- Set as
-
Repeat until convergence is reached.
V. WARM PLASMA CLOSURE (WARMC)
A. Warm plasma closure lattice Boltzmann (WARMC-LB)
Also in this case, we employ the axisymmetric description: the passages that are needed to adapt Eqs. (51), (52), (55), and (56) to this framework are rather lengthy but simple and we report the final full expressions in Appendix A.
VI. NUMERICAL RESULTS
In this section, we present the current capabilities of the method by reproducing known analytical results and we also show side-by-side comparisons between the two closures, highlighting the emergence of momentum spreads anisotropies. We divide the showcase of our results in three different parts: in the first, we start by considering temperature effects in a completely 1D scenario, where our equations of motions are made mono-dimensional by imposing translational symmetry along the radial directions (all derivatives w.r.t. transversal coordinates are, therefore, zero), imposing in Eqs. (33) and (34) (this is equivalent to considering a 1D kinetic momentum space) and considering dimensional tensors in the WARMC case. In this 1D1V set-up, we compare our numerical result with known analytical solutions.30–32,56 Then, we move to the discussion of dispersion relations in a 1D3V setup:57,88 there is still translational invariance along the transversal directions, but a 3D kinetic momentum space is considered [ in Eqs. (33) and (34) and tensors are (3 + 1) dimensional]. Finally, we consider full spatially resolved simulations in a 3D axisymmetric environment (3D3V).
As already mentioned, just like other common PWFA solvers,18 the electromagnetic fields are solved via an FDTD scheme63,64 through numerical integration of the curl Maxwell equations [Faraday's and Ampère's laws in Eqs. (21) and (22)]. It can be shown in fact89 that when these equations are considered together with the continuity equation, the divergence Maxwell counterparts [Gauss laws in Eqs. (21) and (22)] are automatically satisfied, provided that the fields are correctly initialized. For this reason, we properly initialize the electromagnetic fields by solving analytically the Maxwell system with just a rigid Gaussian density (the driving bunch) in its rest-frame and then boost the quantities to the lab-frame.90,91
A. 1D1V results
We adapt our scheme to work in a 1D environment by employing a value of σr which is way bigger than the computational domain along the radial direction Lr so that the driving bunch of Eq. (61) reduces effectively to a single Gaussian profile along the z coordinate, and the system assumes a translational invariance along the r coordinate. Equation (62) is, therefore, employed by imposing . Furthermore, the 1V condition is reached for the LEC by setting in Eqs. (33) and (34), while in the WARMC case is sufficient to initialize to zero the transversal components of the tensor.
We present in Fig. 2 the results for a numerical benchmark of the method obtained by comparing both the two fluid solvers against the analytic solutions that can be obtained by considering Eq. (27) and Eqs. (48)–(52) in a 1D1V environment. The corresponding equations are well studied in the literature: for the LEC, see Sec. VI in Ref. 31 or equivalently;30 for the WARMC, see Ref. 32 and replace the laser driven case with a particle driven case. The theoretical solutions are obtained via finite differences integration of Eq. (6.4) in Ref. 31 (LEC) and Eq. (8) in Ref. 32 (WARMC). Although coming from different set of equations, these solutions are equivalent in the limit of small temperature. Being this the case, they are represented by a single solid black curve in Fig. 2. We observe that both the two methods are perfectly able to reproduce the theoretical results. In this simplified 1D scenario, it is also possible to start appreciating some effects of temperature on the dynamics (the choice of μi is done in order to evidence thermal effects in the first periods of the wave): by comparing against the analytic solutions of the cold fluid model (light purple in Fig. 2), it is possible to see that temperature leads to a decrease in the thermal wavelength of the plasma wave. As it will be seen in Sec. VI B, this effect is in stark contrast to what can be observed in a 3V environment, where the wavelength increases or decreases depending on the selected fluid closure.
B. 1D3V results
C. 3D3V results
We now move our analysis to fully spatially resolved simulations. In Fig. 4, we provide an initial comparison of the two fluid closures in a 3D3V axisymmetric environment. The temperature is set to an initial value of , and the chosen regime of the driving pulse is quasi-linear, . In the figure, we show color plots for four significant quantities: the plasma number density n, the longitudinal fluid velocity uz, the longitudinal accelerating field Ez, and the transversal wakefield . It is possible to appreciate that at variance with the usual dynamics of the cold case,62 the typical wave-like pattern of peaks and valleys is perturbed by the emergence of an acoustic behavior that promotes electron motion out of the radial axis. The presence of pressure waves can be appreciated by the appearance of Mach cone structures95 that can be easily spotted by looking at anyone of the panels of Fig. 4: as the driving bunch travels at the speed of light c, it perturbs the plasma medium. The perturbation propagates at a specific velocity cs (that depends on the initial temperature μi) smaller than c, and the wave-fronts of the perturbation, that in a sub-sonic flow would radially propagate all over the space, are instead constrained in a conical structure around the perturbation: a supersonic Mach cone is observed.
One can then run a set of simulations with different initial temperatures μi and study the inclination of the conical envelope (from which we extract cs) as a function of the temperature. In Fig. 6, we show the results of this analysis that again show good match between the simulations and expectations from theory.
The fluid rest frame pressures and are obtained from the numerics by inverting Eq. (60) (the lab-frame components of the tensor are obtained from the simulations). We note thermal spread anisotropies are mostly relevant in the proximity of the bunch (located at ) where they become of the order of 10% and become less and less important as one proceeds further away from the perturbation. To the best of our knowledge, this is the first time that such analysis is conducted on the WARMC model for a spatially resolved plasma (a study on a laser excited, 1D restricted plasma can be found in Ref. 96) and this preliminary results indicate that the LEC model, that is missing this pressure anisotropy feature by construction, could still be used to characterize plasma behavior in late-stage dynamics studies. We reserve further investigation, in particular studies on the dependency of this anisotropic feature on driving bunch parameters and initial temperatures, for later research.
VII. OUTLOOK AND PERSPECTIVES
In the development of a fluid model for the simulations of PWFA processes, a multitude of physical ingredients has to be taken into account to provide a realistic description of the process. In an earlier paper,62 some of the authors started the exploration of lattice Boltzmann (LB) schemes for the construction of fluid models for PWFA and considered the simplified case of cold fluid models. This paper represents a step forward, in that we have explored LB fluid schemes accounting for thermal effects.30–32,34,39,41,44–46 The inclusion of thermal effects is a rather non trivial task due to the theoretical complication represented by the choice of a proper closure to the set of equations.54 We have handled this problem by selecting two of the most popular closure models that have been discussed in the literature: the first one relies on the assumption of a local equilibrium (LEC),48 while the second one involves truncating the hierarchy of centered equations at an arbitrary order (WARMC).57 We have then shown how to successfully adapt the LB schemes to both closures. If from one side the LEC is nominally not appropriate for a collisionless warm plasma, from the other side, the WARMC is obtained under the assumption of an asymptotically small temperature. Any finite temperature, however small, can raise the question on what is the right closure scheme to obtain the correct fluid model for collisionless warm plasma dynamics. The preliminary comparisons shown in this work, although not presenting strong qualitative differences in the dynamics, give a clear indication that the selection of a closure scheme is pivotal for the quantitative assessment of PWFA experiments. To this aim, a one-to-one comparison between the predictions of fluid models and the PIC simulations (or a numerical solution of the Vlasov equations) could be helpful to shed lights on the matter. Work is in progress along this direction.
The next logical step in the development of the method is the inclusion of ions dynamics. Since plasma ions are way more massive then electrons, their dynamics happens on longer time scales than the ones examined in PWFA studies that target the early stage evolution of system, unless a strongly dense driving bunch is considered.97 There is though a growing research interest for studies that inquire the late time evolution of the system where this kind of dynamics cannot be ignored anymore:36,38,40,98 unlocking the movement of ions brings in a new wealth of physical processes such as soliton dynamic40 and ion channels formation.39 Furthermore, most of these studies invoke thermal effects to explain the acoustic motion of the ions,38,40,41,43 and it is then clear how the development of a method that would be able of handling both temperature and ions dynamics is of the uttermost importance. The inclusion of ion motion in the numerical scheme presented in this paper is theoretically trivial and only brings in a bigger computational effort (more equations need to be integrated): the aforementioned characterization is, therefore, completely within the reach of the LB method.
We would also like to mention a couple of advancements that would make the method presented in this paper a more complete numerical tool for the simulation of wakefield acceleration processes: the first is represented by the possibility of including a laser source field as the plasma perturbating mechanism, as the LWFA has always been an alternative (w.r.t. PWFA) route to wakefield acceleration;8 the second is the possibility of handling non-rigid particle bunches, as the tracking of the driving bunch properties is both an important element of diagnostic and also a feature to be kept under control in modern day PWFA experiments (see, for example, Ref. 99) Again, work in this direction is in progress. On the computational side, we would like to mention that the code developed for this work is amenable to a number of numerical optimizations (such for example porting on GPUs) and it is a firm intention of the authors to realize these improvements in future works.
As the most important design choice presented in this paper is the selection of the LB method as a fluid solver, a few comments are in place. In this paper, we have shown that the LB method is well capable of recovering analytic results, and we would like to mention further advantages that could make LB a convenient and viable option in the context of PWFA simulations. Although PIC methods are able to model the dynamics at the particle level, and hence are able to model kinetic effects, they constantly have to keep at bay their inherent numerical noise.9,100 Therefore, in some situations, it would be preferable to dispose of alternative tools for prototyping and to reserve more complete PIC simulations for later stages. Solvers based on a strict discretization of the Vlasov equations would not suffer of the noise problem and still be able to capture mesoscopic effects, but for any dimensionality more than 1D, they would be too computationally demanding and hungry for memory resources. Fluid solvers (and the LB method among them) present then a viable option for the realization of quick PWFA simulations. They are coarse-grained and hence do not suffer of numerical noise, and still capture a good amount of the physical effects that are encountered in PWFA. Furthermore, the LB method lends itself exceptionally well to parallelization on both CPUs and GPUs.60 In our implementation, we achieved parallelization on multi CPUs using the message passing interface (MPI). This approach involves dividing the computational domain into multiple rectangular sub-domains, corresponding to the number of processors. Leveraging the local nature of computations [Eq. (7), r.h.s.], communications are solely required to exchange populations between neighboring processors during the “streaming” process [Eq. (7), l.h.s.]. By effectively disentangling “compute” and “communicate,” this strategy greatly enhances the parallelization process.101–104 Simulations in this study were conducted on an Intel Xeon E5-2695@2.40 GHz processor. A representative simulation (like those shown in Figs. 4 and 5) run on 96 processors requires approximately 182 min for the LEC-LB model and 368 min for the WARMC-LB model for time steps. Memory requirements for such simulations are ∼1.6 GB for the LEC-LB model and ∼4.3 GB for the WARMC-LB model although these numbers could be decreased by further optimizations of the code. As a final note, we would like to remark that LB takes roughly of the compute time in the simulation, whereas the Maxwell solver for the electromagnetic fields takes 1%. This is to be expected as the amount of computation appearing in LB is significantly higher. Figure 8 presents the execution time per iteration and speedup data for varying numbers of processors (strong scaling).
Furthermore, an advantage of the LB method over Vlasov solvers is that the first adopts a smart pruning of the velocity space,60,61 thus improving the computational efficiency. It is possible to increase the number of discrete velocities Npop (see Sec. II A). However, this leads to a proportional increase in both the computational cost and memory requirements, making it a crucial factor to consider when choosing the LB stencil. Common practice is, indeed, to choose the minimum value Npop that ensures the recovery of the hydrodynamic properties.60 We hasten to remark, however, that at variance with usual hydrodynamic solvers, LB's theoretical formulation is strongly grounded in kinetic theory, and its fluid behavior is obtained via the smart discretization of the velocity space cited above. This is a remarkable feature that makes the method a strong candidate for the inclusion of kinetic effects into PWFA fluid solvers. In fact, recent studies105–107 in other research fields show that LB is, indeed, capable of capturing behaviors beyond hydrodynamics just by increasing the number of discrete kinetic velocities. This could open novel perspectives for developing more refined numerical schemes for simulations of PWFA processes, with the obvious need of a precise comparison/benchmark against some reference PIC/Vlasov simulations. Further investigation on this is in progress.
We finally would like to mention that an alternative route to LB wakefield simulation could be represented by the Relativistic Lattice Boltzmann method.108 This is an extension of LB hydrodynamic schemes (like the ones exposed in Sec. II A) to the theory of special relativity, originally created for simulation in astrophysical109 and condensed matter110,111 contexts, and, therefore, constitutes a promising tool for the simulation of warm plasmas within the LEC assumption. Adapting it to a PWFA framework is though more technical and less immediate w.r.t. the moment matching LB used in this paper, therefore, the authors reserve further development on this line for future works.
ACKNOWLEDGMENTS
The authors gratefully acknowledge Fabio Bonaccorso for his technical support. This work was supported by the Italian Ministry of University and Research (MUR) under the FARE program (No. R2045J8XAW), project “Smart-HEART.” MS gratefully acknowledges the support of the National Center for HPC, Big Data and Quantum Computing, Project CN_00000013 - CUP E83C22003230001, Mission 4 Component 2 Investment 1.4, funded by the European Union - NextGenerationEU.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Daniele Simeoni: Conceptualization (lead); Formal analysis (lead); Investigation (lead); Methodology (equal); Software (equal); Visualization (lead); Writing – original draft (lead). Gianmarco Parise: Conceptualization (supporting); Formal analysis (supporting); Software (equal); Writing – original draft (supporting). Fabio Guglietta: Conceptualization (supporting); Software (supporting); Writing – original draft (supporting). A. R. Rossi: Conceptualization (supporting); Writing – original draft (supporting). James Rosenzweig: Conceptualization (supporting); Writing – original draft (supporting). Alessandro Cianchi: Conceptualization (supporting); Writing – original draft (supporting). Mauro Sbragaglia: Conceptualization (supporting); Methodology (equal); Writing – original draft (supporting).
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.
APPENDIX A: FULL EXPRESSIONS FOR THE FLUID ADVECTION EQUATIONS
- The azimuthal fluid velocity is zero
- Axisymmetric Maxwell equations lead to some electromagnetic field components to be zero
- In the WARMC model, thanks to the strategy explained in the main text (end of Sec. V A) one can derive rest frame quantities